<<

Functional analysis Lecture 4: Normalized functions of ; regular Borel measures; the Riesz–Markov–Kakutani theorem;

Remark 3.10. Assume that the Λ above is positive (i.e. Λf ≥ 0 for f ≥ 0). We observe that the extension Λe is also positive and hence the constructed f is then monotone increasing. Indeed, suppose that there is g ∈ `∞[0, 1] with g(t) ≥ 0 for each t ∈ [0, 1] but Λeg < 0. By normalizing, we can assume that kgk∞ = 1, thus 0 ≤ g(t) ≤ 1 for each t ∈ [0, 1]. Let 1 be the constant 1 function. As Λ is positive and linear, it is monotone (i.e. Λg ≤ Λh if g ≤ h), and hence Λ1 = kΛk. We have k1 − gk∞ ≤ 1 and

Λ(e 1 − g) = Λ1 − Λeg = kΛk − Λeg > kΛk = kΛek, which gives a contradicition. Therefore, positive continuous functionals on C[a, b] correspond to increasing func- tions on [a, b], while general continuous functionals correspond to functions of bounded variation.

Remark 3.11. Since every f ∈ BV([a, b]) is the difference of two increasing functions, we see that every Λ ∈ (C[a, b])∗ can be written as

Λ = Λ+ − Λ−, where Λ+, Λ− are positive. (3.1)

The conclusion of Remark 3.11 can be justified without appealing to Riesz’ Theo- rem 3.9. To this end, let us use a more general language which will be also useful in the sequel. By an ordered vector we mean any E over R equipped with a partial order relation ≥ compatible with the algebraic operations in the sense that x ≥ y implies x + z ≥ y + z for all x, y, z ∈ E and x ≥ y implies λx ≥ λy for all x, y ∈ E, λ ≥ 0. We denote by E+ the positive cone of E defined by {x ∈ E : x ≥ 0}. An ordered vector space which is also a lattice with respect to the given order is called a Riesz space or a vector lattice. If E is a Riesz space and ∨, ∧ denote the lattice operations, then for any vector x ∈ E we define its positive part x+, negative part x− and its modulus |x| by

x+ = x ∨ 0, x− = (−x) ∨ 0, |x| = x ∨ (−x).

We then have x = x+ − x− and |x| = x+ + x−, so this gives a decomposition of x as the difference between two positive elements. A Riesz space is called Archimedean if whenever 0 ≤ nx ≤ y for all n ∈ N and some y ∈ E+, then x = 0. If E is a Riesz space and a normed space with a k·k satisfying the condition

|x| ≤ |y| =⇒ kxk ≤ kyk for all x, y ∈ E, then E is called a normed Riesz space, and in the case where the norm is complete, we call E a Banach lattice. There are plenty of examples of Banach lattices: Rn with the pointwise ordering and any of the norms k·kp (1 ≤ p ≤ ∞), the space C(X) of continuous

1 functions, and Cb(X) of bounded continuous functions on any X, also considered with the pointwise ordering and the supremum norm, as well as Lp(µ)-spaces under the a.e. pointwise ordering. If E is a normed Riesz space, then its dual E∗ can be ordered in a natural way by defining f ≥ g if and only if f(x) ≥ g(x) for every x ∈ E+ (f, g ∈ E∗). We will see (Lemma 3.18 below) that in this way we equip E∗ with a Banach lattice structure1. In particular, the decomposition (3.1) is possible in the Banach lattice (C[a, b])∗. Actually, decomposition (3.1) holds true with a further requirement that

kΛk = kΛ+k + kΛ−k and in a much more general setting of the so-called order unit normed Riesz spaces (see Problem 3.22). For C[a, b] we will obtain that equality with the aid of Hahn’s decom- position theorem. As we know that every continuous linear functional on the real space C[a, b] is rep- resented by a function f ∈ BV[a, b], let us consider the problem of uniqueness of this representation. Obviously, f is not unique as stated in Theorem 3.9, because adding any constant function to f does not change the Riemann–Stieltjes . So, one normalization would be to require that e.g. f(a) = 0. By Jordan’s decomposition, we have f = g −h for some increasing functions g and h, and therefore f has at most countably many points of discontinuity in (a, b). Observe that by modifying the values of f at each of these points, but keeping the variation bounded, we do not change the value of the integral in formula (3.7). Indeed, any g ∈ C[a, b] is Riemann–Stieltjes integrable, hence its integral can be approximated by any of integral sums over partitions with diameters converging to zero2. In particular, we can avoid each of discontinuity of f (except the endpoints) in these partitions. Consequently, another normalization condition which does not affect the validity of (3.7) is that f(t+) = f(t) for each t ∈ (a, b), i.e. f is right- continuous inside the (a, b) (the choice of the left or right limit is arbitrary). Every f ∈ BV([a, b]) satisfying the above two normalization conditions is called a normalized function of bounded variation, and we denote the class of all such functions by

• NBV([a, b]) = {f ∈ BV([a, b]): f(a) = 0 and f is right-continuous in (a, b)}. By modifying the representing function f in Theorem 3.9 as described above, we can guarantee that it belongs to the class NBV([a, b]). Moreover, it turns out that in this class the function f is unique. We will obtain this as a particular case of the ‘uniqueness part’ of Theorem 3.16 below. Before proceeding to that result, let us build a ‘bridge’ be- tween the special Riesz Theorem 3.9 and the much more general Riesz–Markov–Kakutani Theorem 3.16. Given any function f ∈ BV([a, b]), we define a µf on the field F consisting of finite unions of the intervals |y, x](a ≤ y < x ≤ b) which are all right-closed and left-open unless y = a in which case we take the closed interval [a, x]. Namely:

• µf ([a, x]) = f(x) − f(a),

• µf ((y, x]) = f(x) − f(y) for a < y < x ≤ b.

1An interested reader may consult e.g. [C.D. Aliprantis, K.C. Border, Infinite dimensional analysis. A hitchhiker’s guide, 3rd edition, Springer 2006; Ch. 8 and 9]. 2See [W Rudin, Principles of ; Thms. 6.6 and 6.8].

2 Notice that the measure µf is finitely additive and satisfies

n n X [ µf (E) = (f(bj) − f(aj)) for E = |a1, b1] ∪ (aj, bj] ∈ F, j=1 j=2 where a ≤ a1 < b1 ≤ a2 < b2 ≤ ... ≤ an < bn ≤ b. Moreover, µf is bounded because f is of bounded variation. But, as we will see, µf has also another nice property of being regular, provided that f belongs to the class NBV([a, b]).

Definition 3.12. Let F be a field of of a topological space X and µ: F → K be a finitely additive function, where K ∈ {R, C}. We say µ is regular if for every E ∈ F and ε > 0 there exist a set F ∈ F with F ⊆ E and a set G ∈ F with int G ⊇ E satisfying

|µ(C)| < ε for every C ∈ F,C ⊆ G \ F.

If f ∈ NBV([a, b]), it is possible to extend µf to a regular σ-additive measure defined on all Borel subsets of [a, b]. It follows from the following general result on extensions of measures. (Actually, we only need a simplified version of this result, as the measure µf is trivially σ-additive in the sense that there are no countably infinite, but not finite, unions of the intervals |c, d] in F.) Extension theorem. Let F be a field of subsets of some topological space X and let µ: F → K be a bounded, regular, finitely additive set function, where K ∈ {R, C}. Then, µ has a unique regular, σ-additive extension µe: σ(Σ) → K to the σ-algebra σ(F) generated by F. For the proof see e.g. [N. Dunford, J.T. Schwartz, Linear operators (vol. 1: General theory), Wiley–Interscience 1988; Ch. III.5]. This theorem is a combination of two classical measure theory results:

• the Alexandroff theorem which says that every bounded regular finitely additive (real- or complex-valued) set function on a field of subsets of a compact topological space is σ-additive;

• the Carath´eodory extension theorem which guarantees that positive σ-additive mea- sures defined on a ring R can be extended to positive σ-additive measures on the σ-algebra generated by R. (It should also be noted that a measure is regular if and only the positive and negative parts of its real and imaginary parts are regular.) Lemma 3.13. If f ∈ BV([a, b]), then for every t ∈ [a, b) we have

lim V (f, [t, t + ε]) = 0. ε→0+ Proof. (classes) Now, assume that f ∈ NBV([a, b]). Since f is right-continuous on (a, b), we have limε→0 f(t + |ε|) = f(t) for every t ∈ (a, b) and it follows that for every interval |c, d] ∈ F,

n n X Sn o V (f, [c, d]) = sup |µf (Ij)|: Ij ∈ F,Ij ∩ Ik = ∅ for j 6= k, j=1Ij ⊆ |c, d] . j=1

3 (The right-hand side defines the variation of µf on the set |c, d]; we will introduce this notion later on.) Plainly, we can extend this formula to any set E ∈ F in the sense that if E = I1 ∪ ... ∪ Im, with Ij = |cj, dj] ∈ F (1 ≤ j ≤ m), then the left-hand side is Pm j=1 V (f, [aj, bj]), whereas the right-hand side is the variation of µf on E, denoted by |µf |(E). For any such E we can assume that a ≤ c1 < d1 < c2 < d2 < . . . < cm < dm ≤ b. Pick 0 < ε < min (dj − cj), min (cj+1 − dj) 1≤j≤m 1≤j

(The interval (c1, d1 + ε] should be replaced by [a, d1 + ε] in the case where c1 = a.) Then, in view of the formula above and Lemma 3.13, we have

m X |µf |(E \ E1(ε)) = V (f, [cj, cj + ε]) −−−−→ 0 ε→0+ j=1 and m X |µf |(E2(ε) \ E) = V (f, [dj, dj + ε]) −−−−→ 0. ε→0+ j=1

This shows that µf is regular and thus the extension theorem applies. For σ-additive measures we shall use the following common definition of regularity. It does not completely coincide with Definition 3.12 in general, but it does if X is a locally compact, σ-compact3 Hausdorff space. In this case, a positive on X is regular if and only if for every E ⊆ X and each ε > 0 there exist a closed set F and an V such that F ⊆ E ⊆ V and µ(V \ F ) < ε (classes).

Definition 3.14. Let µ be a positive Borel measure on a locally compact Hausdorff space X. A Borel set E ⊆ X is called outer regular (resp. inner regular) if

µ(E) = inf µ(V ): E ⊆ V,V is open

(resp. µ(E) = sup µ(K): K ⊆ E,K is compact ). The measure µ is called regular if every Borel of X is both outer and inner regular.

We have proved that any f ∈ NBV([a, b]) gives rise to a bounded, regular, finitely additive measure on F to which we can apply the extension theorem. Also, it is not difficult to see that, in the converse direction, the regularity of µf implies that f is right- continuous on (a, b). Summarizing, we have the following result.

Proposition 3.15. Let f ∈ BV([a, b]) (real- or complex-valued), f(a) = 0, let F be the field of finite unions of the intervals of the form [a, d] and (c, d] (a < c < d ≤ b), and let µf : F → R be the finitely additive (real- or complex-valued) measure associated with f as above. Then, µf is regular if and only if f ∈ NBV([a, b]) in which case µf can be uniquely extended to a Borel, regular, σ-additive measure.

3X is σ-compact if it is a countable union of compact sets.

4 In the Riesz–Markov–Kakutani theorem we deal with positive functionals on the space of compactly supported continuous functions. We use the following notation:

• supp(f) = {x ∈ X : f(x) 6= 0}

• Cc(X) = {f : X → K | supp(f) is compact} Also, when given a topological space X, we write:

• K ≺ f if K ⊆ X is compact, f ∈ Cc(X), 0 ≤ f ≤ 1 and f|K = 1,

• f ≺ V if V ⊆ X is open, f ∈ Cc(X), 0 ≤ f ≤ 1 and supp(f) ⊂ V ,

• K ≺ f ≺ V if K ≺ f and f ≺ V . Two basic topological tools which are essential in the proof are the following results: Urysohn’s lemma. Let X be a locally compact Hausdorff space, V ⊆ X be an open set and K ⊂ V a compact set. Then, there exists a function f ∈ Cc(X) such that K ≺ f ≺ V.

Partition of unity. Let X be a locally compact Hausdorff space. Suppose V1,...,Vn ⊆ X are open set and K is a compact set, K ⊂ V1 ∪...∪Vn. Then, there exist functions hi ≺ Vi (for 1 ≤ i ≤ n) satisfying

h1(x) + ... + hn(x) = 1 for every x ∈ K.

In what follows, the space Cc(X) can consist of real- or complex-valued continuous functions with compact . Correspondingly, Λ can be real-valued and R-linear, as well as complex-valued and C-linear. However, the crucial assumption is that Λ is positive. Notice that the measure µ produced by the theorem below can be infinite. For k R +∞ example, one can think of X = R (or, more generally R ) and Λf = −∞ f(x) dx in which case the measure µ is the on R. Theorem 3.16 (Riesz–Markov–Kakutani). Let X be a locally compact Hausdorff space and Λ be a positive linear functional on Cc(X). Then, there exist: a σ-algebra M in X which contains all Borel subsets of X, and a positive measure µ on M satisfying the following conditions: Z (a)Λ f = f dµ for every f ∈ Cc(X); X (b) for every compact set K ⊂ X we have µ(K) < ∞; (c) for every E ∈ M, µ(E) = inf µ(V ): E ⊆ V,V is open ;

(d) for every open set E and every E ∈ M with µ(E) < ∞,

µ(E) = sup µ(K): K ⊆ E,E is compact ;

(e)( X, M, µ) is complete, i.e. if E ∈ M, µ(E) = 0 and A ⊂ E, then A ∈ M. Moreover, the measure µ is unique in the class of positive measures on M satisfying conditions (a)–(d).

5 Proof. The proof consists of three main steps. First, we show that there is at most one measure µ satisfying the announced properties. Next, we provide definitions of µ and M, together with an auxiliary class MF . We do it in such a way that assertion (c) follows from the very definition, while (e) is trivial. Finally, and this will be the toughest part, we shall show that M is a σ-algebra containing all Borel set, µ is σ-additive on M and satisfies (a), (b) and (d). For clarity, the last, most complicated step will be split into ten parts. Uniqueness. If M and µ satisfy (c) and (d), then µ is completely determined by its values on compact sets. Hence, if ν is another positive measure on M satisfying (a)–(d), it is enough to show that µ(K) = ν(K) for every compact set K ⊂ X. For any ε > 0, conditions (b) and (c) imply that there exists an open set V such that K ⊂ V and ν(V ) < ν(K)+ε. By Urysohn’s lemma, there is a function f with K ≺ f ≺ V , and hence Z Z Z Z µ(K) = 1K dµ ≤ f dµ = Λf = f dν ≤ 1V dν = ν(V ) < ν(K) + ε. X X X X Therefore, µ(K) ≤ ν(K) and, by symmetry, µ(K) = ν(K). Construction of M and µ. We define:

µ(V ) = sup Λf : f ≺ V for any open set V ⊆ X, (3.2)

µ(E) = inf µ(V ): E ⊆ V,V is open for an arbitrary E ⊆ X. (3.3) Notice these two definitions are compatible. For, if E is open, then for any open set U ⊇ E, the values of µ defined by (3.2) satisfy µ(E) ≤ µ(U). Hence, µ(E) defined by formula (3.3) is the same as that defined by (3.2). Next, define MF to be the collection of all sets E ⊂ X such that

µ(E) < ∞ and µ(E) = sup µ(K): K ⊆ E,K is compact .

Finally, let  M = E ⊂ X : E ∩ K ∈ MF for every compact set K ⊂ X . Observe that by formula (3.2) we have guaranteed condition (c). Of course, the function µ is monotone, i.e. A ⊆ B implies µ(A) ≤ µ(B) and it implies that condition (e) is also satisfied. Proof of properties (a), (b) and (d).

∞ Part 1. µ is an outer measure, that is, for any sequence (Ei)i=1 of subsets of X we have ∞ ∞  [  X µ Ei ≤ µ(Ei). i=1 i=1

First, we show this inequality for two open sets. So, assume V1,V2 ⊆ X are open and pick a function g with g ≺ V1 ∪ V2. Using the partition of unity, we produce functions h1, h2 such that h1 ≺ V1, h2 ≺ V2 and h1(x) + h2(x) = 1 for each x ∈ supp(g). Hence, h1g ≺ V1, h2g ≺ V2 and g = h1g + h2g which yields

Λg = Λ(h1g) + Λ(h2g) ≤ µ(V1) + µ(V2).

6 Passing to supremum over all the functions g with g ≺ V1 ∪ V2, we obtain µ(V1 ∪ V2) ≤ µ(V1) + µ(V2). The rest of the proof is left for the classes.

Part 2. For any compact set K ⊂ X we have K ∈ MF and µ(K) = inf Λf : K ≺ f .

Consider any function f with K ≺ f and α ∈ (0, 1). Define Vα = {x ∈ X : f(x) > α} and note that K ⊂ Vα and αg ≤ f provided that g ≺ Vα. Hence,  −1 µ(K) ≤ µ(Vα) = sup Λg : g ≺ Vα ≤ α Λf. Passing to the limit as α → 1, we get µ(K) ≤ Λf < ∞. Therefore, K satisfies the requirements of the definition of MF , thus K ∈ MF . We have established assertion (b). Part 3. Every open set V ⊆ X satisfies µ(V ) = sup µ(K): K ⊆ V,K is compact , and hence MF contains all open sets V for which µ(V ) < ∞. (classes)

∞ Part 4. Let (Ei)i=1 be a sequence of mutually from the class MF . Then,

∞ ∞  [  X µ Ei = µ(Ei). i=1 i=1 S∞ S∞ Moreover, if µ( i=1Ei) < ∞, then i=1Ei ∈ MF . We will prove that

µ(K1 ∪ K2) = µ(K1) + µ(K2) for compact K1,K2 ⊂ X,K1 ∩ K2 = ∅. (3.4)

Fix any ε > 0. By Urysohn’s lemma, there is a function f ∈ Cc(X) such that f|K1 = 1 and f|K2 = 0. In view of Part 2, we can pick a function g with

K1 ∪ K2 ≺ g and Λg < µ(K1 ∪ K2) + ε.

Notice that K1 ≺ fg and K2 ≺ (1 − f)g. Since in Part 2 we have proved that µ(K) ≤ Λh for any compact K and any h with K ≺ h, we have

µ(K1) + µ(K2) ≤ Λ(fg) + Λ(g − fg) = Λg < µ(K1 ∪ K2) + ε. Therefore, we obtain formula (3.4) by the fact that ε > 0 was arbitrary. The rest of the proof is left for the classes.

Part 5. For every E ∈ MF and any ε > 0 there exist a compact set K and an open set V such that K ⊆ E ⊆ V and µ(V \ K) < ε. (classes)

Part 6. If A, B ∈ MF , then A ∪ B,A ∩ B,A \ B ∈ MF . (classes) TBC

7